首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Herein we present an unusual phosphine-free method to fabricate water soluble CdSeS nanocrystals in cubic structure. In this method, glycerin was used as a stabilizing agent replacing tri-n-octylphosphine oxide (TOPO). Water solution of Na2SeO3 in polyethylene glycol was utilized as Se source. 3-Mercaptopropionic acid (MPA) provides S source. The phosphine-free Se and S sources were found to be highly reactive and suitable for the synthesis of CdSeS nanocrystals. XRD and HRTEM images confirm the formation of CdSeS nanocrystals in zinc blende structure. The absorption peaks on UV-vis spectra of as-prepared CdSeS nanocrystals are tunable from 330 nm to 440 nm, which blue shifts to shorter wavelength side in comparison with that of pure CdSe nanocrystals. The cubic CdSeS nanocrystals demonstrate narrow PL emissions spectra between 464 and 615 nm. Transmission electron microscopy images show the uniformity for the size distribution of the ternary QDs. Series water soluble CdSe1–xSx (x = 0∼1) nanocrystals have also been synthesized using Na2SeO3 and Na2S solution as the Se-S co-sources. Tunable band gap energies of CdSe1–xSx (x = 0∼1) nanocrystals upon chemical composition x have been achieved, the gap ranges from 290 nm to 558 nm.  相似文献   

2.
Abstract

Using CD measurements we show that the interaction of netropsin to poly(dA-dT)·poly(dA-dT) involves two binding modes at low ionic strength. The first and second binding modes are distinguished by a defined shift of the CD maximum and the presence of characteristic isodichroic points in the long wavelength range from 313 nm to 325 nm. The first binding mode is independent of ionic strength and is primarily determined by specific interaction to dA·dT base pairs. Employing a netropsin derivative and different salt conditions it is demonstrated that ionic contacts are essential for the second binding mode. Other alternating duplexes and natural DNA also exhibit more or less a second step in the interaction with netropsin observable at high ratio of ligand per nucleotide. The second binding mode is absent for poly(dA)·poly(dT). The presence of a two-step binding mechanism is also demonstrated in the complex formation of poly(dA-dT)·poly(dA-dT) with the distamycin analog consisting of pentamethylpyrrolecarboxamide. While the binding mode I of netropsin is identical with its localization in the minor groove, for binding mode II we consider two alternative interpretations.  相似文献   

3.
Enhancement of intensity and wavelength tunability of emission are desirable features for light-emitting device applications. We report on the large and tunable blue shift (60 nm) in emission from an environment-sensitive fluorophore (Coumarin153) embedded in Ag plasmonic random media. Coumarin 153 having emission at 555 nm, show a systematic blue shift (to 542, 503 and 495 nm) upon infiltration into random media fabricated by Ag nanowires of different aspect ratio (hence, surface plasmon resonances at 426, 445 and 464 nm). The blue shift is due to the fast dynamic surface-enhanced fluorescence mechanism and can be tuned by controlling the surface plasmon resonance and hotspot density in random media. Enhanced emission at desired wavelength is achieved by using nanostructures having higher extinction coefficient but same-surface plasmon resonance. Ag nanostructures of different aspect ratio used for fabricating the random media are synthesized by chemical route.  相似文献   

4.
CdSe nanoparticles were synthesized by green route and chemical route methods. In the green route method the samples were capped by starch and in the chemical route method the samples were capped by mercaptoacetic acid (MAA). The samples were characterized by powder X‐ ray diffraction (XRD) and transmission electron microscopy (TEM). Both the samples showed zinc blend structure. The optical absorption spectra and Fourier transform infrared (FTIR) spectra were also studied. A blue shift was seen in the absorption spectra as compared with the bulk as well as the sample capped by starch. TEM images showed agglomeration for the starch‐capped sample as compared with the MAA‐capped sample. The particle size for the sample capped by MAA was found to be less as compared with the starch‐capped sample. A blue shift in the photoluminescence (PL) spectra was also recorded for the samples prepared by the chemical route as compared with the sample prepared by the green route as well as the bulk. The PL peak shifted towards the red side and increase in the peak intensity occurred with the change in the excitation wavelength. Change in PL intensity was observed with different pH at 685 nm. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

5.
Surfactants are widely used in the purification and research of structure and function of the protein complexes in photosynthetic membrane. To elucidate the mechanism of interaction between surfactants and photosystem Ⅰ (PSⅠ), effects of two typical surfactants, Triton X-100 and sodium dodecyl sulfate (SDS) on PSⅠ, were studied at different concentrations. The results were: SDS led to the reduction of apparent absorption intensity and blue shift of absorption peaks; while Triton X-100 led to the decrease of apparent absorption intensity in red region and blue shift of the peak, but to the increase of apparent absorption intensity in blue region. The fourth derivative spectra show that the longwavelength (669 nm and 683 nm) absorbing chlorophyll a was affected greatly and their relative changes of absorbance were axially symmetrical. The presence of surfactant could make the long wavelength fluorescence emission decrease greatly and a new fluorescence peak appeared around 680 nm, it was obvious that the surfactant interceded the transfer of excitation energy from antenna pigments to reaction center. The surfactants might affect the microenvironment of proteins, even the structure of PSⅠ protein subunits and hence changed the binding status of pigments with protein subunits, or the pigments might be released from the subunits. All of these might affect the absorption and the transfer of excitation energy.  相似文献   

6.
Using CD measurements we show that the interaction of netropsin to poly(dA-dT).poly(dA-dT) involves two binding modes at low ionic strength. The first and second binding modes are distinguished by a defined shift of the CD maximum and the presence of characteristic isodichroic points in the long wavelength range from 313 nm to 325 nm. The first binding mode is independent of ionic strength and is primarily determined by specific interaction to dA.dT base pairs. Employing a netropsin derivative and different salt conditions it is demonstrated that ionic contacts are essential for the second binding mode. Other alternating duplexes and natural DNA also exhibit more or less a second step in the interaction with netropsin observable at high ratio of ligand per nucleotide. The second binding mode is absent for poly(dA).poly(dT). The presence of a two-step binding mechanism is also demonstrated in the complex formation of poly(dA-dT).poly(dA-dT) with the distamycin analog consisting of pentamethylpyrrolecarboxamide. While the binding mode I of netropsin is identical with its localization in the minor groove, for binding mode II we consider two alternative interpretations.  相似文献   

7.
CdSe/ZnS quantum dots (QDs) exhibited fluorescence emission blue shifts when conjugated to antibodies or DNA aptamers that are bound to bacteria. The intensity of the shifted emission peak increased with the number of bound bacteria. Curiously, the emission was consistently shifted to approximately 440-460 nm, which is distinctly different from the major component of the natural fluorescence spectrum of these QDs. This minor emission peak can grow upon conjugation to antibodies or aptamers and subsequent binding to bacterial cell surfaces. We hypothesize that the wavelength shift is due to changes in the chemical environment of the QD conjugates when they encounter the bacterial surface and may be due to physical deformation of the QD that changes the quantum confinement state. Regardless of the mechanism, these remarkable emission wavelength shifts of greater than 140 nm in some cases strongly suggest new applications for QD-receptor conjugates.  相似文献   

8.
A novel blue luminescent 6‐chloro‐2‐(4‐cynophenyl) substituted diphenyl quinoline (Cl‐CN DPQ) organic phosphor has been synthesized by the acid‐catalyzed Friedlander reaction and then characterized to confirm structural, optical and thermal properties. Structural properties of Cl‐CN‐DPQ were analyzed by Fourier transform infrared (FTIR), nuclear magnetic resonance (NMR) spectroscopy, X‐ray diffraction technique (XRD) and scanning electron microscopy (SEM) and energy dispersive analysis of X‐ray (EDAX) spectroscopy. FTIR spectra confirmed the presence of different functional groups and bond stretching. 1H–NMR and 13C–NMR confirmed the formation of an organic Cl‐CN‐DPQ compound. X‐ray diffraction study provided its crystalline nature. The surface morphology of Cl‐CN‐DPQ was analyzed by SEM, while EDAX spectroscopy revealed the elemental analysis. Differential thermal analysis (TGA/DTA) disclosed its thermal stability up to 250°C. The optical properties of Cl‐CN‐DPQ were investigated by UV–vis absorption and photoluminescence (PL) measurements. Cl‐CN‐DPQ exhibits intense blue emission at 434 nm in a solid‐state crystalline powder with CIE co‐ordinates (0.157, 0.027), when excited at 373 nm. Cl‐CN‐DPQ shows remarkable Stokes shift in the range 14800–5100 cm?1, which is the characteristic feature of intense light emission. A narrow full width at half‐maximum (FWHM) value of PL spectra in the range 42–48 nm was observed. Oscillator strength, energy band gap, quantum yield, and fluorescence energy yield were also examined using UV–vis absorption and photoluminescence spectra. These results prove its applications towards developing organic luminescence devices and displays, organic phosphor‐based solar cells and displays, organic lasers, chemical sensors and many more.  相似文献   

9.
Taylor's blue (1,9-dimethylmethylene blue, DMMB+) associates with DNA, at least in part, through intercalation as is evidenced from the red shift in the absorption maximum, diminution of the fluorescence, and induced circular dichroism in the presence of nucleic acid. Irradiation of DMMB+/covalently closed circular supercoiled phiX174 phage DNA complex at lambda > 520 nm leads to DNA nicking in a dose-dependent manner.  相似文献   

10.
One of the most significant advances in nanoscience and nanotechnology was partially driven by plasmonic effect of some noble metal nanostructures with different shapes and sizes. By controlling the geometry of metal nanostructures, their surface plasmon resonance (SPR) peaks could be tuned from the visible to the near-infrared region with various applications in sensors, optoelectronic, nanomedicine, and specifically cancer therapy. In this study, we have prepared gold nanoboxes (NBs) using the galvanic replacement between Ag nanocubes (NCs) and aqueous gold solution. Ultraviolet visible (UVvis) spectroscopy, scanning electron microscopy (SEM), X-ray diffraction (XRD), and transmitting electron microscopy (TEM) were used to characterize silver NCs and gold NBs. The primary silver NCs were synthesized by conventional polyol method at the presence of sodium sulfide to highly tune the shape and size of the NCs. Optimized cubic silver nanostructures were obtained at 90 μl of sodium sulfide injection into the solution. Moreover, the effect of quality of the cubic structure on the shape and uniformity of gold NBs was investigated. Gold NBs with hollow interior structure and SPR peak ranging from 480 to 800 nm were successfully obtained at different injection volumes of HAuCl4 into the solution. It was demonstrated that increasing the volume of HAuCl4 solution to about 3 mL can increase the pore number and size until the primary structure collapses into small pieces. It was also found that the concentration of gold NBs and the corresponding SPR peak intensities decrease due to pore size enhancement and decline of charge density on the surface of metal hollow nanostructures.  相似文献   

11.
The photoluminescence (PL) characteristics of semiconductor CdSe quantum dots (QDs) aggregated on Cu nanowires (NWs) were studied in detail. The PL relaxation dynamic data show that Cu NWs improve the PL intensity of CdSe QDs by accelerating the emission relaxation rate. The temperature‐dependent PL data and excitation intensity‐dependent PL data suggest that the activation energy of CdSe QDs might decrease due to the excellent heat transfer properties and the plasmon effect of Cu NWs. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

12.
B.G. De Grooth  J. Amesz 《BBA》1977,462(2):247-258
An analysis was made of the changes of pigment absorption upon illumination of chromatophores of Rhodopseudomonas sphaeroides at ?35 °C, described in the preceding paper (de Grooth, B. G. and Amesz, J. (1977) Biochim. Biophys. Acta 462, 237–246). Comparison of the light-induced difference spectra in the carotenoid region obtained without additions, and in the presence of N-methylphenazonium methosulphate and ascorbate as donor-acceptor system showed that the latter spectrum was not only about 10 times larger in amplitude, but also red-shifted with respect to the first one. Together with the shape of the difference spectrum, this indicated that the spectrum obtained in the presence of a donor-acceptor system is due to an electrochromic shift of the absorption spectrum of a carotenoid by a few nm towards longer wavelength, caused by a delocalized potential across the chromatophore membrane. The results of an analysis of the kinetics of the absorbance changes near the zero points of the spectrum were in quantitative agreement with the extent of the red shift and indicated a shift of 0.25 nm for a single electron transfer per reaction center, and shifts of up to 4 nm when the electron transport is stimulated by a donor-acceptor system. For bacteriochlorophyll B-850 the shift is three times smaller.Analysis of the overall absorption spectrum showed that there are at least two pools of carotenoid. The carotenoid that shows electrochromism has absorption bands at 452, 481 and 515 nm, and comprises about one-third of the total carotenoid present; the remaining pool absorbs at about 7 nm shorter wavelength and does not show an electrochromic response to illumination. Both pools presumably consist of spheroidene; the differences in band location may be explained by the assumption that only the first pool is subjected to a local electric field which induces an electric dipole even at zero membrane potential. Similar results were obtained at room temperature and with a mutant of Rps. sphaeroides (G1C)-containing neurosporene.  相似文献   

13.
In this work, CdSe quantum dots (QDs) were synthesized by a simple and rapid microwave activated approach using CdSO4, Na2SeO3 as precursors and thioglycolic acid (TGA) as capping agent molecule. A novel photochemical approach was introduced for the growth of CdS QDs and this approach was used to grow a CdS shell around CdSe cores for the formation of a CdSe/CdS core–shell structure. The core–shells were structurally verified using X‐ray diffraction, transmission electron microscopy and FTIR (Fourier‐transform infrared (FTIR)) spectroscopy. The optical properties of the samples were examined by means of UV–Vis and photoluminescence (PL) spectroscopy. It was found that CdS QDs emit a broad band white luminescence between 400 to 700 nm with a peak located at about 510 nm. CdSe QDs emission contained a broad band resulting from trap states between 450 to 800 nm with a peak located at 600 nm. After CdS shell growth, trap states emission was considerably quenched and a near band edge emission was appeared about 480 nm. Optical studies revealed that the core–shell QDs possess strong ultraviolet (UV) ? visible light photocatalytic activity. CdSe/CdS core–shell QDs, showed an enhancement in photodegradation of Methyl orange (MO) compared with CdSe QDs.  相似文献   

14.
The interactions between 2‐mercaptoethanol, dimercaprol and CdSe quantum dots (QDs) in organic media have been investigated by spectral methods. The results showed that the fluorescence (FL) emission of CdSe QDs gradually decreased, with a slight red‐shift, after adding thiols to CdSe QDs solutions. With the increase of the concentrations of thiols, the resonance light scattering (RLS) signal of CdSe QDs had been strongly enhanced in the wavelength range 300–500 nm, which was confirmed by the formation of larger CdSe QDs particles. The effect of thiols on the FL emission of CdSe QDs could be described by a Stern–Volmer‐type equation with the concentration ranges 1.0 × 10–6–7.5 × 10–4 mol/L for 2‐mercaptoethanol and 1.0 × 10–7–2.5 × 10–5 mol/L for dimercaprol. The possible mechanism of the interaction was proposed according to the results of UV‐vis absorption and micro‐Raman spectroscopy. The results indicated that FL quenching was mainly attributable to the exchange of the QDs surface molecules. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

15.
We have designed a simple, one‐step synthesis of CdSe quantum dots with photoluminescence frequencies ranging from the red through to the violet region of the solar spectrum. The photoluminescence peaks have FWHM of 30 nm indicating absorption over a narrow range of wavelengths. The effect of solvent type and solvent boiling point on the physical and photoluminescence properties of the quantum dots has been studied. High boiling point, non‐polar solvents shift the photoluminescence peak to longer wavelengths and low boiling point, polar solvents shift the photoluminescence peak to shorter wavelengths. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
The interaction of the radioprotector 1-methyl-2-[2-(methylthio)-2-piperidinovinyl]quinolinium iodide (VQ) with linear and supercoiled pIBI30 DNA was studied by flow linear dichroism spectroscopy, equilibrium dialysis, circular dichroism, and UV absorption spectroscopy. The negative linear dichroism spectra of VQ-DNA complexes throughout the 220-500 nm wavelength region, a red shift in the VQ main absorption band (at 452 nm) of 1-2 nm upon binding to DNA, and a concentration-dependent unwinding of supercoiled DNA suggest that the primary mode of interaction of VQ with DNA (at least at low concentrations) is intercalative in nature. A least-squares analysis of the equilibrium dialysis binding of VQ to supercoiled DNA using the McGhee-von Hippel equation gives an association constant K = 7300 +/- 300 M-1, and an exclusion number n in the range of 3.3-5.3. The lower value of n is obtained when effects of polyelectrolytes are also taken into account. Because quinolinium iodide derivatives with different substituents and DNA binding affinities can be synthesized, this family of compounds could be employed to probe relationships, if any, between radioprotective efficacy and DNA binding affinity.  相似文献   

17.
Fasick JI  Applebury ML  Oprian DD 《Biochemistry》2002,41(21):6860-6865
The wild-type mouse ultraviolet (UV) and bovine blue cone visual pigments have absorption maxima of 358 and 438 nm, respectively, while sharing 87% amino acid identity. To determine the molecular basis underlying the 80 nm spectral shift between these pigments, we selected several amino acids in helices II and III for site-directed mutagenesis. These amino acids included: (1) those that differ between mouse UV and bovine blue; (2) the conserved counterion, Glu113; and (3) Ser90, which is involved in wavelength modulation in avian short-wavelength sensitive cone pigments. These studies resulted in the identification of a single amino acid substitution at position 86 responsible for the majority of the spectral shift between the mouse UV and bovine blue cone pigments. This is the first time that this amino acid by itself has been shown to play a major role in the spectral tuning of the SWS1 cone pigments. A single amino acid substitution appears to be the dominant factor by which the majority of mammalian short-wavelength sensitive cone pigments have shifted their absorption maxima from the UV to the visible regions of the spectrum. Studies investigating the role of the conserved counterion Glu113 suggest that the bovine and mouse SWS1 pigments result from a protonated and unprotonated Schiff base chromophore, respectively.  相似文献   

18.
Monodispersed ZnS nanoparticles have been successfully synthesized by a chemical precipitation method in an air atmosphere using polyvinylpyrrolidone (PVP) and sodium hexametaphosphate (SHMP) as surfactants. The synthesized nanoparticles were characterized by X‐ray diffraction (XRD), transmission electron microscopy (TEM), Fourier transform infrared spectrometer (FT‐IR), UV–Vis optical absorption and photoluminescence (PL) spectra. Prepared surfactants capped ZnS nanoparticles are highly homogeneous and well dispersed. Optical absorption spectra showed a strong blue shift from the uncapped particles due to the quantum confinement effect. The capped ZnS emission intensity is enhanced than more the uncapped particles. The size of the synthesized particles is around 4–6.5 nm range. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

19.
Three A–π–A or D–π–D perylene diimide ( PDI ) derivatives with varied groups on π-conjugate were synthesized and characterized. The photophysical properties of these compounds were systematically studied by spectral experiments and density functional theory calculations. All compounds displayed intense absorption bands at 300–800 nm wavelengths. However, diverse groups on the π-conjugate influenced the UV–vis absorption. Electron-withdrawing groups on PDI-2 caused a slight red shift at the 350–400 nm wavelength and a blue shift after 400 nm wavelength. At the same time, the electron-donating substituents on PDI-3 caused an obvious red shift of this band. These PDI derivatives exhibited emission in solution at room temperature (λem = 500–850 nm). The quantum yield of PDI-3 decreased, while the electron-donating substituents were introduced to the π-conjugated motifs. However, the quantum yield of PDI-2 increased when electron-withdrawing substituents were introduced to the π-conjugated motifs. In addition, PDI-1 and PDI-2 exhibited broad triplet transient absorption in the visible region. These photophysical properties could help us to understand the relationship between structure and photophysical properties of perylene diimide derivatives and exploit more original perylene diimide-based optical functional materials.  相似文献   

20.
Photolysis of E-[ring-2-14C]urocanic acid (UA) with native or denatured calf thymus DNA leads to covalent binding of the radiolabel to the nucleic acid. A similar observation is made upon photolysis of the labeled UA with the polyribonucleotides, in which case a strong preference is observed for binding to poly[U]. DNA or poly[U], which had been reacted with UA and purified by dialysis and multiple precipitations, releases UA upon further irradiation with 254 nm light (as expected for cyclobutane adducts). Quantum efficiencies for binding of the UA to native DNA have been measured at 308 and 266 nm and are 0.30 x 10(-5) and 1.3 x 10(-4), respectively, at comparable reactant concentrations. The large increase at the shorter wavelength (where DNA absorption is more competitive) is taken as evidence for the primary role of a DNA excited state in initiating the binding reaction(s).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号